Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Article https://doi.org/10.

1038/s41467-023-36524-x

Microwave-assisted design of nanoporous


graphene membrane for ultrafast and
switchable organic solvent nanofiltration
Received: 31 August 2022 Junhyeok Kang1, Yeongnam Ko2, Jeong Pil Kim1, Ju Yeon Kim1, Jiwon Kim1,
Ohchan Kwon1, Ki Chul Kim 2 & Dae Woo Kim 1
Accepted: 2 February 2023

Layered two-dimensional materials can potentially be utilized for organic


Check for updates solvent nanofiltration (OSN) membrane fabrication owing to their precise
1234567890():,;
1234567890():,;

molecular sieving by the interlayer structure and excellent stability in harsh


conditions. Nevertheless, the extensive tortuosity of nanochannels and bulky
solvent molecules impede rapid permeability. Herein, nanoporous graphene
(NG) with a high density of sp2 carbon domain was synthesized via sequential
thermal pore activation of graphene oxide (GO) and microwave-assisted
reduction. Due to the smooth sp2 carbon domain surfaces and dense nano-
pores, the microwave-treated nanoporous graphene membrane exhibited
ultrafast organic solvent permeance (e.g., IPA: 2278 LMH/bar) with excellent
stability under practical cross-flow conditions. Furthermore, the membrane
molecular weight cut-off (MWCO) is switchable from 500 Da size of molecule
to sub-nanometer-size molecules depending on the solvent type, and this
switching occurs spontaneously with solvent change. These properties indi-
cate feasibility of multiple (both binary and ternary) organic mixture separa-
tion using a single membrane. The nanochannel structure effect on solvent
transport is also investigated using computation calculations.

Chemical purification is a crucial process in various industries, such (CON), porous organic cages (POC), metal-organic frameworks
as petrochemicals and pharmaceuticals, to purify valuable products (MOF), and functionalized graphene5–8.
and recycle solvents or catalysts1. Conventional purification pro- Graphene and its derivatives are the most widely utilized two-
cesses based on distillation and crystallization are costly and inap- dimensional (2D) materials for membrane fabrication, which has
plicable for heat-sensitive products2. Alternatively, organic solvent been widely applied to separation applications such as gas separa-
nanofiltration (OSN) membranes capable of separating molecules tion, water treatment, organic solvent purification, pervaporation,
between 100–2000 Da in size are promising owing to their envir- and nanofiltration9–15. Particularly, graphene oxide (GO) is effective
onmentally friendly operation, high energy efficiency, and integrated for membrane fabrication because large-scale fabrication is feasible
system3,4. Specifically, when the membrane has switchable molecular via practical continuous coating methods (bar or slot-die coating)
separation properties by external stimuli or chemical conditions, the and its narrow interlayer spacing enables precise molecular
OSN-based separation process can be further integrated and made separation11,16,17. However, narrow interlayer spacing and lengthy
cost-effective because most practical chemical purification pro- diffusion pathways typically result in limited organic solvent flow18.
cesses aim to separate multiple mixed molecules. As a proof of Therefore, various approaches have been reported to modify nano-
concept, few membranes with switchable molecular separation channel structure, such as pore generation, flake size control, and
properties have been demonstrated using covalent organic networks hybridization with other nanomaterials, which can decrease solvent

1
Department of Chemical and Biomolecular Engineering, Yonsei University, Yonsei-ro 50, Seodaemun-gu Seoul (03722), Republic of Korea. 2Department of
Chemical Engineering, Konkuk University, Seoul 05029, Republic of Korea. e-mail: audw1105@yonsei.ac.kr

Nature Communications | (2023)14:901 1


Article https://doi.org/10.1038/s41467-023-36524-x

diffusion length or increase nanochannel size18–20. While these wave treatment can generate denser nanopores during sp3 carbon
methods can produce OSN membrane with rapid solvent permea- rearrangement or decomposition, further enhancing solvent permea-
tion, switchable molecular separation phenomena have been rarely tion. Notably, previous chemical and thermal reduction methods uti-
reported and the majority of membranes have been utilized to lized for graphene membranes solely remove oxygen-containing
separate single or binary mixtures18,21. groups with partial recovery of sp2 domains because recrystallization
Herein, nanoporous graphene with rich sp2 carbon domains were of amorphous carbons require extremely high temperatures above
prepared for OSN membrane fabrication. Dense nanopores on the 2000 °C25–27. Furthermore, as thermally activated nanoporous gra-
basal plane of graphene were activated by spontaneous thermal phene is extremely defective, conventional thermal and chemical
annealing of GO. Subsequently, microwave treatment was conducted reduction methods are inapplicable. In contrast, the microwave
to increase the proportion of sp2 carbon domains by transforming treatment is effective at recrystallizing the sp3 carbon structure by
defective sp3 carbons at elevated local temperatures. Graphitic struc- increasing the local temperature of graphene owing to its explosive
tural changes owing to thermal activation and microwave treatment carrier scattering under microwave irradiation28.
were systemically investigated. The membrane OSN performances The prepared MNG powder was dispersed in 1-methyl-2-
were investigated using a dead-end and cross-flow system. Further- pyrrolidone (NMP) via sonication. After removing large particles and
more, molecular separation properties of microwave-treated nano- unexfoliated flakes by centrifugation, the MNG solution was vacuum
porous graphene (MNG) membrane were demonstrated to be filtered on nylon support. Figure 1b illustrates the nanochannel
switchable in response to the solvents employed to separate multiple structures of MNG membrane. The dense nanopore can effectively
mixtures of subnanometer-sized organic molecules through repetitive provide additional flow channels and filter large molecules. Owing to
solvent exchange and filtration using a single membrane. sp2 domain recovery after microwave treatment, the MNG membrane
can possess a narrower interlayer spacing than the NG membrane, and
Results the remaining sp3 carbons serve as spacers to support stacked gra-
Fabrication of microwave-treated NG membrane phene sheets. In addition, the nanochannel size of the MNG membrane
Figure 1a depicts the overall fabrication procedure for MNG mem- is spontaneously variable based on the swelling degree of organic
branes. First, to prepare nanoporous graphene (NG), the as-prepared solvents within the MNG layer, resulting in switchable molecular
GO powder was thermally annealed at 200 °C for 5 min under ambient separation phenomena, which is subsequently described in detail.
conditions, which activates numerous nanopores on the graphene Figure 2a, b depicts the SEM images of MNG particles and exfo-
surface by decomposing oxygen-containing groups22. Thermal liated nanosheet, respectively. As-prepared GO sheets were densely
annealing is effective at producing dense nanopores; however, oxygen stacked, whereas the NG layers were expanded during thermal acti-
group decomposition frequently results in amorphous sp3 carbon vation by CO and CO2 gas emissions (Supplementary Fig. 1)18. The
formation. Because the sp3 carbon structure may impede solvent expanded graphene layers maintained a hierarchical structure after
molecule transport owing to steric hindrance23,24, NG was treated with microwave treatment, but frequent porous defects were observed on
microwave to recover sp2 domains on the graphene surface25. Voiry the graphene surface, which can be attributed to the decomposition of
et al. reported that high-quality graphene composed with a high pro- oxygen-containing groups and sp3 carbons caused by microwave-
portion of sp2 carbons can be prepared via treating reduced graphene induced local high temperatures (Fig. 2a). Figure 2b illustrates that
oxide with microwave25, while microwave treatment effects on nano- the exfoliated MNG flake exhibits a lateral size of several hundred
porous graphene have been not investigated. In addition, the micro- nanometers and a thickness of ~6 nm, revealed using atomic force

Fig. 1 | Microwave-assisted design of nanoporous graphene membrane. a Preparation of sp2 carbon enriched nanoporous multilayer graphene membrane enabled by
microwave-assisted reduction. b Membrane working principle for switchable molecular separation in response to solvent types.

Nature Communications | (2023)14:901 2


Article https://doi.org/10.1038/s41467-023-36524-x

Fig. 2 | Structure of sp2-enriched nanoporous graphene and its membrane. and corresponding EDS mapping images. The numbers represent the atomic
a, b SEM images of MNG particle and exfoliated nanosheet. c AFM image of MNG weight percentages. h, i Photographic images of MNG dispersion in NMP at 0.5 mg/
nanosheet and its height profile. d, e HR-TEM images of MNG. f Corresponding mL of concentration and MNG membrane. j, k Top-view and cross-sectional SEM
selected area electron diffraction (SAED) pattern of MNG obtained from d. g TEM images of MNG membrane on nylon support.

microscopy (AFM) and corresponding height profile (Fig. 2c). Strong structure of graphene, indicating the recovery of sp2 carbon bonds
inter-sheet connections via van der Waals interaction or generation of from amorphous sp3 carbons following microwave treatment. Sharp
intersheet chemical bonding appear to impede single-layer exfoliation hexagonal patterns in the corresponding SAED pattern of MNG cor-
of MNG sheets, which were mostly produced as multilayered roborate the microwave-induced transformation of sp2 domains
sheets29,30. In optimizing microwave strength, more defects were (Fig. 2f). However, the SAED pattern is not as sharp as that of graphite
observed to appear on the graphene plane with increasing microwave or single-layer graphene grown by chemical vapor deposition owing to
power (Supplementary Fig. 2) and 700 W was selected as the optimal dense nanopore formation31. Energy dispersive spectroscopy (EDS)
condition to generate dense nanopores and restore sp2 domains. The mapping images for carbon and oxygen (Fig. 2g) further demonstrate
treatment duration was maintained within a few seconds to preserve that oxygen groups were effectively eliminated by microwave treat-
the graphene structure, which would otherwise burn. ment, as evidenced by the low oxygen quantities (3%).
As illustrated in Fig. 2d–f, high-resolution TEM images, and cor- Figure 2h depicts a photograph of the MNG solution dispersed in
responding SAED and atomic mapping images were acquired to NMP at a concentration of 0.5 mg/mL using sonication and cen-
directly visualize the crystalline structure of the MNG flake. Although trifugation. The MNG solution was filtered on commercial nylon sup-
GO possessed numerous oxidized regions with partial sp2 regions, the ports (200 nm pore size) to fabricate membranes (Fig. 2i). Because of
oxidized regions were eliminated after thermal activation (Supple- the high degree of reduction, the membrane color was as dark as
mentary Fig. 3). However, because the activation temperature (200 °C) graphite paper. Figure 2j, k displays the SEM images of uniform MNG
is too low to induce recrystallization of sp3 carbon, amorphous sp3 layers with 200 nm thickness. Defective structures such as cracks and
regions were observed and low crystallinity was confirmed by the aggregated particles were undetected. In addition, the membrane was
broad hexagonal SAED pattern in Supplementary Fig. 3. On the other flexible and delamination of graphene layer was not observed as
hand, the TEM images of MNG (Fig. 2d) show abundant nanopores and shown in the bending test (Supplementary Fig. 4), indicating good
the high-resolution TEM image (Fig. 2e) reveals a hexagonal lattice adhesion between graphene and polymer support.

Nature Communications | (2023)14:901 3


Article https://doi.org/10.1038/s41467-023-36524-x

Fig. 3 | Chemical structure of graphene materials. a XPS survey scan, and b corresponding atomic percentage of GO, NG, MNG, CrGO, and MCrGO. c XPS C1s spectra and
d Raman spectra of GO, NG, MNG, CrGO, and MCrGO. Raman spectra were measured using an excitation wavelength of 532 nm.

Effect of pore structure on the degree of microwave reduction further validates the substantial reduction degree of graphene with
XPS and Raman spectroscopy investigation were conducted to analyze microwave treatment (Supplementary Fig. 8), indicating the sup-
the effects of pore activation and microwave treatment on the crys- pressed thermal decomposition of graphenes after microwave treat-
talline structure of nanoporous graphene (Fig. 3). First, the chemical ment. The contact angles of solvents (Supplementary Fig. 9) exhibited
structures of graphene materials were investigated using XPS as enhanced hydrophobic and oleophilic properties of the MNG mem-
depicted in Fig. 3a–c and Supplementary Fig. 5. The XPS survey scan brane than graphene oxide owing to the low proportion of oxygen-
and corresponding atomic percentage of samples (Fig. 3a, b) exhibit a containing groups.
reduced oxygen atomic percentage from 30.2% for GO to 7.9% for Raman spectroscopy was employed to investigate the crystalline
MNG due to the sequential thermal and microwave treatments. Che- structure (Fig. 3d) because the presence of sharp D, G, and 2D bands
mically reduced graphene oxide (CrGO) with hydrazine exhibited a straightforwardly indicates the presence of long-range ordered sp2
reduced oxygen atomic percentage of 14.2%, which further decreased carbon domains32. While directly visualized TEM images in Fig. 2 con-
to 5.3% after microwave treatment (MCrGO, microwave-treated che- firm the presence of ordered carbon crystals, the relatively large beam
mically reduced graphene oxide). MNG has a slightly greater oxygen size (several micrometers in diameter) of Raman spectroscopy is more
atomic percentage than MCrGO because NG comprised of a defective informative and reveals the crystal structure on a large scale. The D-
porous structure that is difficult to recrystallize. In addition, defective band/G-band intensity ratio decreased from 1.02 for GO to 0.94 for NG
graphene is well known to exhibit low electrical conductivity, which owing to the decomposition of oxygen-containing groups following
may decrease the local temperature of graphene under microwave thermal activation. The broad D-band width of NG indicates amor-
irradiation28. phous sp3 carbon formation33. However, the Raman spectrum of MNG
The chemical bonds of graphene were further investigated as (shown in green) reveals a sharp D-band and the existence of a 2D peak,
depicted in XPS C1s spectra in Fig. 3c and Supplementary Figure 5. GO indicating sp2 carbon recovery by microwave treatment25. Interest-
exhibits oxygen-containing groups at C–O (285.9 eV), C = O (287.5 eV), ingly, the D/G intensity ratio of MNG significantly increased to 1.57
and COOH (289.8 eV). Due to the decomposition of oxygen-containing owing to dense nanopore formation and high concentration of carbon
groups, the C–O peak intensity significantly decreased after the first edges in nanopores34,35. The spectra shapes are identical to that of
thermal annealing. The dominant C–C peak of MNG, generated from a nanoporous graphene produced by etching CVD graphene with
defective amorphous carbon structure, was observed after microwave plasma or ion beam exposure35. The D-band intensity increased with
treatment. While the C1s spectrum shape of CrGO is similar to that of increasing microwave power, possibly owing to the enhanced gen-
MNG, CrGO, and MCrGO primarily exhibited a C–C peak and an eration of defective pores (Supplementary Fig. 10). The D/G intensity
additional C–N (284.6 eV) peak originated from hydrazine. As shown in ratio of MNG is remarkably higher and the 2D peak intensity is lower
Fig. 3a, the atomic nuclei scans reveal low amounts of oxygen- than that of MCrGO, indicating that the NG pore structure hinders
containing groups in MCrGO, reconfirming a higher degree of reduc- carbon crystallization compared to CrGO36. Because GO is unreactive
tion than MNG. The reduction of oxygen-containing groups was fur- towards microwave owing to its insulation properties, a comparison
ther confirmed by the FT-IR spectra (Supplementary Figs. 6–7), with neat GO was not performed. Moreover, the average size of sp2
revealing the significantly decreased peak intensities related to domains was estimated using the Raman spectra of the samples
oxygen-containing groups after thermal activation and microwave (Supplementary Fig. 11)37. The sp2 crystallite size of MNG was calcu-
treatment. As depicted in Supplementary Fig. 6b, the peak intensities lated to be 13.5 nm, which is less than the size of GO (19.1 nm) and NG
decreased with increasing microwave power owing to the enhanced (20.7), implying that the sp2 domain size decreased throughout the
decomposition of oxygen-containing groups at higher local tempera- thermal annealing and microwave treatment, despite the increasing
tures. The low weight loss of MNG and MCrGO during the TGA test density of sp2 domains.

Nature Communications | (2023)14:901 4


Article https://doi.org/10.1038/s41467-023-36524-x

Fig. 4 | Interlayer structure of graphene materials. a Normalized XRD patterns of state and after swelling in a good solvent. d–g Cross-sectional HR-TEM images of
GO, NG, MNG, CrGO, and MCrGO. b XRD patterns of MNG membrane after NG (d, e) and MNG (f, g), respectively, and corresponding FFT images.
immersion in various solvents. c Schematic structure of MNG membrane in dry-

Nanopore generation was also confirmed by N2 adsorption– MNG (~3.5 Å) reduced below that of NG (3.6 Å) owing to enhanced
desorption isotherms and corresponding pore size distribution (PSD) reduction and increased crystallinity, and the d-spacing remained
by Barrett–Joyner–Halenda (BJH) model (Supplementary Fig. 12 and identical regardless of microwave power (Supplementary Fig. 13). Due
Supplementary Table 1). Significant N2 adsorption by NG at relative to the presence of both sp2 regions and sp3 regions in the basal plane of
pressures <0.1 indicates the presence of micropores38. MNG adsorbs MNG, a diffraction peak of MNG is considered to be composed of two
more N2 than NG, and the PSD reveals MNG pore size to be <3 nm. patterns: defective region stacking and pi-pi stacking of sp2 regions39.
Furthermore, the Brunauer–Emmett–Teller (BET) surface area This assumption is corroborated by the MCrGO peak, which exhibits a
increased from 306 m2/g for NG to 419 m2/g for MNG. These results less broad and right-shifted pattern because MCrGO is primarily
indicate additional nanopore formation by microwave treatment. structured with sp2 carbons resulting from a higher degree of reduc-
Based on the preceding observations, it is concluded that nanoporous tion and the absence of dense nanopores.
graphene with high crystallinity can be successfully prepared using The XRD patterns in solvents were further analyzed to determine
sequential thermal activation and microwave treatment, notwith- the interlayer spacing variations in the MNG membrane in response to
standing the presence of a small proportion of sp3 carbon or oxygen. solvent type (Fig. 4b). The d-spacing values of the MNG membrane are
similar in water, ethanol, and IPA, which are known to be poor solvents
Interlayer structure and swelling behavior in solvents for graphene dispersion, and these values are near to that of the MNG
X-ray diffraction (XRD) patterns were obtained from graphene films to membrane in the dry state. In contrast, the d-spacing value was sig-
elucidate the interlayer structure of graphene after thermal activation nificantly increased in NMP which is a well-known good solvent for
and microwave treatment, as depicted in Fig. 4a. The d-spacing graphene dispersion40. The interlayer spacing variation in solvents is
reduced from 8.3 Å for GO to 3.6 Å for NG owing to the removal of illustrated in Fig. 4c. The MNG membrane exhibits a combined struc-
oxygen-containing groups after thermal activation18. The NG diffrac- ture with defective sp3-stacked regions and graphitic sp2-stacked
tion peak is broader than that of graphite owing to the less-ordered regions. Particularly, the spacing composed of defective sp3 regions is
structure of NG resulting from the presence of amorphous sp3 carbons more easily intercalated by solvent molecules than the narrow spacing
on the graphene plane18. After microwave treatment, the d-spacing of of sp2 regions. Therefore, the interlayers may further swell in good

Nature Communications | (2023)14:901 5


Article https://doi.org/10.1038/s41467-023-36524-x

Fig. 5 | Organic solvent nanofiltration performance of MNG membrane. blue (MnB), rhodamine B (RhB), acid red 1 (AR), congo red (CR), methyl blue (MB),
a Permeance of pure organic solvents across 200 nm thick NG and MNG mem- brilliant blue G (BBG), and evans blue (EB) at 3 bar: IPA (b), ethanol (c), NMP (d).
branes as a function of viscosity. b–d Nanofiltration performance of MNG mem- e Comparison of IPA nanofiltration performances. All data points were averaged
branes tested with several organic solvents and dye molecules including methylene from at least three samples.

solvents. In contrast, the interlayer structure with sp2-stacked regions solvents. The permeance of various pure organic solvents was initially
could be rigidly maintained owing to the pi–pi interactions of sp2 evaluated (Fig. 5a and Supplementary Table 2) at 3 bar using the
regions, which is crucial for fabricating stable membranes in solvents. 200 nm thick NG and MNG membranes (Fig. 2k and Supplementary
As depicted in Fig. 4d–g, cross-sectional HR-TEM images of the Fig. 14). The MNG membrane demonstrated ultrafast permeances for
MNG membrane were obtained to reveal its interlayer structure. Both various organic solvents, ranging from 13117 LMH/bar for acetone to
samples were prepared using the focused ion beam technique. The 1300 LMH/bar for 1-butanol, which are surprisingly near to the solvent
TEM images of both NG and MNG membranes depict the stacked permeances of a bare nylon support. In summary, the solvent per-
graphene layers; however, the MNG membrane exhibits better aligned meance was inversely proportional to the solvent viscosity. The sol-
and compactly stacked graphene layers. The enhanced ordered vent permeance in 2D lamellar membranes can be calculated
structure of the MNG membrane was further confirmed by the FFT theoretically by using classical Hagen-Posieuile (HP) equation41. The
patterns denoted by yellow arrows, indicating a broader arc pattern permeance of MNG membrane (2278 LMH/bar of IPA) is exceptionally
from the NG sample with larger d-spacing. higher than the theoretical values as shown in Supplementary Fig. 15. A
possible reason can be the smoother surface of highly reduced nano-
OSN performance of MNG membrane in various solvents porous graphene with dense sp2 carbon regions, making the transport
The membrane was tested as depicted in Fig. 5 to estimate the nano- of ordered solvent molecules easier in the nanochannels. In addition,
filtration performance of the MNG membrane in various organic low friction at graphene surface by low oxygen groups and sp2 regions

Nature Communications | (2023)14:901 6


Article https://doi.org/10.1038/s41467-023-36524-x

can be favorable for the solvent transport, as observed in the fast water Moreover, any crack on the membrane surface was not observed after
transport through supported graphene nanochannel41,42. the filtration test even though a dye cake was formed due to the fast
Remarkably, toluene and water exhibited relatively low per- solvent permeance (Supplementary Fig. 24). In addition, the mem-
meance. The low permeance of toluene is expected to arise from the brane showed excellent diafiltration performance with binary mixed
large molecular size of toluene, resulting in hindered transport within dye mixture feed under cross-flow conditions, maintaining high
the interlayers43. Moreover, a lower water flux than other solvents rejection rates of EB above >99%, while Methyl orange (MO) rejection
could result from the hydrophobic surface of MNG, which is unfa- rate was ~0% (Supplementary Fig. 25).
vorable to water sorption. Several previous studies had either expec-
ted the selective permeation of alcohol or nonpolar solvent through Multiple mixture separation by MNG membrane
hydrophobic nanoporous graphene channels44, or the promotion of To explore the feasibility of separation for mixed molecules, binary
hydrogen bonding with water by residual oxygen-containing groups, mixture separation tests were conducted using the MNG membrane
thereby hindering water molecule diffusion45,46. In addition, it is under a dead-end condition (Fig. 6 and Supplementary Fig. 26). The
emphasized that the MNG membrane exhibited significantly higher test was conducted at 3 bar with a total concentration of 1000 ppm in
permeances than the NG membrane owing to the increased produc- IPA for the binary feed solution. The feed solution was filtered through
tion of nanopores and the smoother surface comprised of dense sp2 the MNG membrane, following which the retentate was re-dispersed
carbon regions. Detailed mechanisms of enhanced organic solvent and re-filtrated. The process was repeated several times to perfectly
transport in graphene nanochannel are subsequently discussed herein separate the binary mixture (Fig. 6a). After initial filtering, the yellow
using computational simulations. solution (MO, 327 Da) was permeated without the dark red solution
Figure 5b–d illustrates the OSN performance of the MNG mem- (AR, 509 Da), and the re-dispersed retentate solution darkened as the
brane in IPA, ethanol, and NMP with probe dye molecules ranging in yellow solution of MO was completely extracted (Fig. 6b). In the UV–Vis
size from 300 to 1000 Da. The tests were conducted at 3 bar and spectra (Fig. 6c–e and Supplementary Fig. 27), while the mixed feed
10 mg/L of dye concentration (Supplementary Fig. 16). As depicted in solution contains two absorbance peaks of AR and MO (Fig. 6c), the
Fig. 5b, the MNG membrane exhibited nearly 100% high rejection rates permeate solutions exhibited a single peak of smaller MO by the fifth
for EB (961 Da), CR (697 Da), and AR (509 Da), whereas smaller mole- cycle (Fig. 6e). Furthermore, the re-dispersed retentate solution only
cules exhibited lower rejection rates such as RhB (479 Da, 50.5%) and exhibited a peak corresponding to larger AR after the fifth cycle, indi-
MnB (320 Da, 20.7%). The NG membrane demonstrated lower rejec- cating the successful separation of most smaller MO molecules and the
tion rates than the MNG membrane below 600 Da, such as AR (84.6%), concentration of larger AR molecules based on size exclusion using the
RhB (31.5%), and MnB (18.1%), indicating the enhanced rejection per- MNG membrane (Fig. 6d). Figure 6f, g exhibits the purity and con-
formance by the better ordered and narrower interlayers of MNG. The centration of AR and MO in feed and permeate side after each filtration
MNG membrane was evaluated in ethanol and NMP under identical cycle, indicating the separation of highly concentrated AR and MO with
conditions (Fig. 5c, d). Interestingly, the MWCO of the MNG membrane a high purity of nearly 100% on each side after five cycles of filtration.
increased from 509 Da to 800 Da in ethanol; furthermore, the MNG Most importantly, due to the ultrafast solvent permeance of the MNG
membrane exhibited low rejection rates below 20% for all dyes in NMP. membrane, a single filtration of 100 mL of solution was completed
Because the membrane exhibited stability in NMP, membrane damage within a minute, indicating that the operation efficiency can be sig-
cannot account for the low rejection rate (Supplementary Fig. 17). nificantly greater than conventional OSN membranes or distillation.
According to the XRD patterns of swollen MNG membranes, the low A membrane with switchable molecular separation properties is
rejection rates in ethanol and NMP may be attributable to the enlarged verily required to purify multiple mixtures with single equipment
interlayer spacing in ethanol and NMP (Fig. 4b). Moreover, the smaller because separation processes can be accomplished without
effective/solvated diameters of dyes in ethanol can cause low rejection membrane replacement. To demonstrate the switchable molecular
rates47. These results indicate that the separation performance of MNG separation property of the MNG membrane, the filtration test was
membrane can be altered in response to different solvent types. In conducted at 3 bar with a feed of ternary dye mixtures containing MO
addition to the aforementioned solvents, we also tested MNG mem- (327 Da), AR (509 Da), and EB (961 Da) in IPA (Fig. 7) and the separation
branes with acetone and methanol, which are also one of the most was conducted by solvent exchange to exploit the solvent-dependent
widely used solvents in organic chemistry (Supplementary Fig. 18). switchable interlayer structure of MNG (Fig. 7a). The brown feed
MWCO was around 450 Da in acetone and around 750 Da in methanol. solution exhibits three absorbance peaks of MO, AR, and EB as shown
We also confirmed that the retentate was concentrated after filtration in Fig. 7b. After the initial filtration, the yellow solution (MO) was
due to the filtered dyes on feed side, indicating that the rejection permeated with only the MO absorbance peak as MO is below the
occurred by molecular sieving rather than adsorption (Supplemen- 509 Da of MWCO in IPA. During the second filtration, the pink solution
tary Fig. 19). (AR) was permeated with an AR peak as AR is below the 800 Da of
The OSN performance of the MNG membrane was compared to MWCO in ethanol. Finally, the blue solution (EB) was completely per-
that of prior OSN membranes using the upper bound based on pure meated in NMP with an EB peak, demonstrating that the ternary
IPA permeance and MWCO (Fig. 5e and Supplementary Table 3). The molecular mixtures were successfully separated by solvent exchange
MNG membrane exhibited an ultrafast pure IPA permeance of 2278 with the MNG membrane. Furthermore, the filtration process was
LMH/bar and the MWCO of 509 Da, which significantly exceeded the repeated with the ternary mixture feed solution to confirm the rever-
upper bound of previously reported OSN membranes. In addition, the sible switching properties. The MNG membrane exhibited the same
MNG membrane performance in ethanol was compared to that of result after the second test as depicted in the UV–Visible absorbance
other OSN membranes (Supplementary Fig. 20 and Supplementary spectra and photo images of permeate solutions. The cycle test con-
Table 4) and the MNG membrane stability was verified using an firmed that the interlayer structure of the MNG membrane is switch-
immersion test and long-term filtration test under cross-flow condi- able based on the solvents and that the separation performance is
tions (Supplementary Figs. 21–25). For 30 days, the MNG membrane reversible. For polymeric membranes, the changes of MWCO by
proved extremely stable in various solvents including water, IPA, acid, swelling of polymer in organic solvents have been widely reported.
and alkaline solutions (Supplementary Fig. 21). Furthermore, the However, the MWCO changes of polymeric membranes are normally
membrane exhibited extremely stable rejection rates (>99%) for two not fast and not precise as we are reporting with our graphene mem-
days during the cross-flow filtration test (Supplementary Figs. 22–25), brane. We want to emphasize that the switchable separation property
which is achieved by strong pi–pi interaction between MNG flakes. of MNG is attributable to the fast interlayer structure modulation of

Nature Communications | (2023)14:901 7


Article https://doi.org/10.1038/s41467-023-36524-x

Fig. 6 | MNG membrane for binary dye mixture separation under dead-end and permeate solutions after each filtration cycle. c UV–Visible absorbance spectra
filtration. The test was conducted at 3 bar and dilution was conducted with 100 mL of AR, MO, and their mixture solutions in IPA. d, e UV–Visible absorbance spectra of
of IPA. The total concentration was 1000 ppm in IPA and the weight ratio of the re-dispersed retentate and permeate solutions after each filtration cycle. f, g Purity
dyes was 1:1. a Schematic illustration of binary dye mixture separation. The colored and concentration of AR and MO in the re-dispersed retentate and permeate
square and circle are dye molecules. b Photographs of feed, re-dispersed retentate, solutions after each filtration cycle.

graphene by solvents. Indeed, the solvent exchange during the test As depicted in the snapshots, the solvent molecule permeability
was conducted immediately after filtering solvent, so this phenom- strongly depends on the solvent identity and interlayer spacing. Water
enon is more close to switching of MWCO by solvent type while solvent and toluene tend to cluster at a d-spacing of 6.4 Å, exhibiting multiple
swelling in the graphene channel is the origin of the switching distributions of locally overlapping graphene layers. This can be
phenomena. attributed to the Connolly-analyzed diameters of water (3.37 Å) and
A steered MD simulation approach was employed to investigate toluene (3.83 Å) molecules, which are larger than the accessible
the change in the thermodynamic free energy, i.e., the potential of interlayer spacing (3.0 Å) calculated under the condition of the van der
mean force, as a solvent permeates through the graphene layers. This Waals diameter (3.4 Å) of graphene carbon (Supplementary Fig. 28).
thermodynamic quantity represents the solvent molecule perme- Therefore, the graphene layers swell to create a local space for the
ability within the graphene layers. Figure 8a depicts snapshots per- solvent molecules, while the graphene layers locally overlap to com-
taining to the observed behaviors of solvent molecules and graphene pensate for the swollen space. The Connolly-analyzed diameter of
layers during the steered MD simulations. The model utilized herein is toluene is assumed to be identical to methane because aromatic rings
based on a pristine graphene layer. While graphene oxide or reduced of toluene molecules are positioned parallel to the graphene layers.
graphene oxide possess oxygen-containing groups or defective sp3 Using identical reasoning, the solvent molecules become well dis-
carbon regions, which are known to affect solvent permeation through tributed throughout entire graphene layers as the layer spacing
interaction or steric hindrance23,24, it is straightforward to understand increases at d-spacings of 7.4, 8.4, and 13.4 Å, which correspond to the
intrinsic solvent permeation through the graphene channel when accessible interlayer spacings of 4.0, 5.0, and 10.0 Å, respectively
pristine graphene spacing is considered. In addition, the effect of (Supplementary Fig. 28). No local overlapping of the graphene layers is
nanopores is disregarded because nanopores with dimensions of observed at these interlayer spacings. Consequently, these solvents
several nanometers are sufficiently large to permit solvent permeation, have significantly higher potentials of mean force at the layer spacing
and the primary resistive channel is the narrow interlayer spacing of of 6.4 Å compared to larger interlayer spacings, as shown in Fig. 8b,
stacked graphene. indicating the highest resistance against the solvent permeation at the

Nature Communications | (2023)14:901 8


Article https://doi.org/10.1038/s41467-023-36524-x

Fig. 7 | Switching MNG membrane performance for ternary dye mixture total concentration was 10 ppm in IPA and the weight ratio of the dyes was 1:1:1. The
separation under dead-end filtration. a Schematic image of ternary dye mixture test was conducted at 3 bar. The feed solvent was IPA. Ethanol and NMP were used
separation. The colored square, circle, and triangle are dye molecules. b UV–Visible for second and third filtration respectively.
absorbance spectra and photographic images of feed and permeate solutions. The

layer spacing of 6.4 Å. Using the same reasoning, the minimum layer demonstrate that the aforementioned difference in the potentials of
spacing at which ethanol molecules homogeneously pass through non- mean force would result from the difference in the interaction energy,
overlapping graphene layers is determined to be 8.4 Å, considering which follows the order of toluene < water < ethanol. Note that a larger
the Connolly-analyzed diameter (4.79 Å) of ethanol. Therefore, the negative interaction energy value results in stronger interaction
potentials of mean force at the interlayer spacings of 6.4 and 7.4 Å are between graphene layers and solvent molecules. Consequently, the
significantly greater compared with those at layer spacings of 8.4 and van der Waals-based strongest interaction between graphenes and
13.4 Å. Conclusively, these results combinedly imply that the perme- aromatic rings in toluene molecules leads to the strongest resistance
ability of solvent molecules would be significantly enhanced at inter- against the solvent permeation, whereas stronger hydrogen bonding
layer spacings with a free volume that can accommodate solvent characteristics in water molecules would result in a stronger resistance
molecules without any swelling behaviors of graphene layers, as indi- of water relative to ethanol against solvent permeation. The simulation
cated in green in Fig. 8a. results imply that a larger interlayer spacing than the size of the solvent
Further comparison of potentials of mean force at the interlayer is critical to achieving high solvent permeation through graphene
spacing of 13.4 Å for all three types of solvents reveals that toluene has nanochannels, and that permeation can be further enhanced by
the highest resistance to solvent permeation, while ethanol exhibits modifying the graphene surface with low interaction energy while
the least resistance (Fig. 8c). As depicted in Fig. 8d, the simulated maintaining the same interlayer spacing. It is expected that the tur-
interaction energies between graphene layers and solvent molecules bostratic nanochannel of the proposed graphene membrane is one of

Nature Communications | (2023)14:901 9


Article https://doi.org/10.1038/s41467-023-36524-x

Fig. 8 | Steered MD simulation of solvent permeation in graphene layers. a. c Potentials of mean force of water, toluene, and ethanol at a d-spacing of 13.4 Å.
a Snapshots of solvents (water, toluene, and ethanol) passing through graphene d Interaction energy between graphene layers and solvents within the d-spacing
layers with various d-spacings. b Potentials of mean force displayed as a function of of 13.4 Å.
the center of mass of solvent molecules for the steered MD simulations described in

the reasons for an ultrafast solvent flux, despite the fact that the nanoporous graphene membrane. Using graphene as a building
swelling degree of different solvents is generally uncontrollable in block, a nanoporous multilayer graphene membrane with ultrafast
experiments, unlike the simulation model. solvent permeation and switchable rejection performance was
developed. By recovering the sp2 carbon structure from sp3 carbons
Discussion through microwave treatment, an enhanced ordered interlayer
This study demonstrates the easy fabrication of nanoporous gra- structure with dense nanopores can be produced. Consequently, the
phene with high crystallinity by combining thermal activation and membrane exhibited flexible and narrow MWCO owing to the swel-
microwave treatment. Previous reduction methods, such as chemical ling degree of interlayers depending on the solvent type. Further-
reduction, are unable to restore the hexagonal lattice of defective more, the membrane was demonstrated to feasibly perform the
and oxidized graphene. However, microwave treatment is highly separation of multiple molecular mixtures, which can be utilized for
effective for regenerating hexagonal carbon lattice as maintaining chemical extraction and concentrations in various application fields,
nanoporous structure. In addition, the treatment process only including petrochemicals and pharmaceuticals. Further, the pro-
requires a few seconds, which is desirable for the future scalability of posed graphene preparation methods can be applied to sensors,

Nature Communications | (2023)14:901 10


Article https://doi.org/10.1038/s41467-023-36524-x

energy storage devices, adsorbents, and electrocatalysis, all of which (298 K and 1 atm) ensembles for 5 ns to have graphene layers with a
require a high surface area and electrical conductivity. targeted layer spacing in which space between graphene layers was
filled with solvent molecules. Four distinctive layer spacings, namely
Methods 6.4, 7.4, 8.4, and 13.4 Å, were introduced for each solvent case to
Synthesis of NG powder identify suitable layer spacings for the facilitated solvent permeation.
GO powder was prepared using a modified Hummers’ method as Nose–Hoover thermostat and barostat were used to control the tem-
described in our previous studies48. First, 2 g of graphite powder was perature and pressure under the condition of the NPT ensemble,
dissolved in 150 mL of sulfuric acid. Second, 7 g of potassium per- respectively. All these simulations were performed using the LAMMPS
manganate (KMnO4) was slowly added to the graphite dispersion and simulation package with the DREIDING force field49,50. Next, a steered
stirred at 35 °C for 5 h in a water bath. After stirring, 200 mL of deio- MD simulation approach implemented in the LAMMPS simulation
nized (DI) water was slowly added to the dispersion in an ice bath. package was performed to predict free energy changes associated with
Subsequently, 100 mL of hydrogen peroxide (H2O2) was poured into the behaviors of solvent molecules passing through the graphene
the solution. The products (GO) were washed and filtered several times layers. Note that the steered MD simulation approach is suitable to
with DI water to remove the remaining chemicals. Finally, GO powder understand the thermodynamic preference of the designated path-
was obtained by freeze-drying. The prepared GO powder was loaded ways of solvent molecules. During the steered MD simulations, a spring
into a quartz tube, which was then placed into a furnace set to 200 °C. constant of 1000 kcal/mol/Å and pulling velocity of 10−5 Å/fs were
The GO powder was thermally annealed in the furnace at 200 °C under applied to solvent molecules in the simulation models while graphene
ambient conditions for 5 min to activate nanopores. The activated layers were not presumed to be pulled out. The free energy profiles of
powder, NG, was collected after cooling to room temperature in air. the solvent molecules passing through the space between graphene
layers during the steered MD simulations were recorded in form of the
Production of MNG membrane potential of mean force (PMF). All the steered MD simulations were
The as-prepared NG powder was placed into a commercial microwave performed under NPT (298 K and 1 atm) ensemble for 1 ns.
oven. The NG powder was treated at 700 W for 2 s. The resultant MNG
powder was collected and dissolved at a concentration of 0.5 mg/mL in Data availability
NMP. The solution was dispersed using a horn-sonicator (Sonic VCS The data supporting the findings of this study are available within the
505) in an ice bath for 300 min. The obtained MNG solution was cen- paper and Supplementary Information. Any additional detail can be
trifuged at 5400 × g (7000 rpm) for 30 min to remove large and requested from the corresponding authors.
undispersed particles, and the dispersed MNG solution was collected
from the supernatant. The as-prepared MNG solution was diluted with References
deionized water and subsequently filtered through nylon support 1. Szekely, G., Jimenez-Solomon, M. F., Marchetti, P., Kim, J. F. &
(Whatman, 0.2 µm pore size). After filtration, the MNG membranes were Livingston, A. G. Sustainability assessment of organic solvent
dried in an oven at 60 °C overnight to eliminate any leftover solvents. nanofiltration: from fabrication to application. Green. Chem. 16,
4440–4473 (2014).
Binary dye mixture separation using a dead-end system 2. Marchetti, P., Jimenez Solomon, M. F., Szekely, G. & Livingston, A. G.
100 mL of feed solution containing methyl orange (MO; 327.33 g·mol−1) Molecular separation with organic solvent nanofiltration: a critical
and acid red 1 (AR; 509.43 g·mol−1) dyes (1:1 weight ratio) at a total review. Chem. Rev. 114, 10735–10806 (2014).
concentration of 1000 mg/L in IPA was filtered through an MNG 3. Nie, L., Chuah, C. Y., Bae, T. H. & Lee, J. M. Graphene‐based
membrane using a homemade dead-end cell until the entire filtration advanced membrane applications in organic solvent nanofiltration.
of the solvent. Next, 100 mL of pure IPA was introduced to the dead- Adv. Funct. Mater. 31, 2006949 (2020).
end cell to re-disperse the retentate, and the solution was filtered 4. Kim, J. F. et al. In situ solvent recovery by organic solvent nanofil-
again. The process was repeated 12 times. The concentrations of feed, tration. ACS Sustain. Chem. Eng. 2, 2371–2379 (2014).
permeate, and re-dispersed retentate solutions were calculated using 5. Liu, J. et al. Smart covalent organic networks (CONs) with “on-off-
UV–Vis spectroscopy. on” light-switchable pores for molecular separation. Sci. Adv. 6,
eabb3188 (2020).
Ternary dye mixture separation using a dead-end system 6. He, A. et al. A smart and responsive crystalline porous organic cage
Under the dead-end condition, 100 mL of feed solution containing membrane with switchable pore apertures for graded molecular
methyl orange (MO; 327.33 g·mol−1), acid red 1 (AR; 509.43 g·mol−1), sieving. Nat. Mater. 21, 463–470 (2022).
and Evans blue (EB; 960.8 g·mol−1) dyes (1:1:1 weight ratio) at the total 7. Jeong, G. Y. et al. Metal-organic framework patterns and mem-
concentration of 10 mg/L in IPA was filtered through the MNG mem- branes with heterogeneous pores for flow-assisted switchable
brane. After the completion of solvent filtration, 100 mL of pure separations. Nat. Commun. 9, 3968 (2018).
ethanol was added to the dead-end cell to re-disperse the retentate. 8. Liu, J. et al. Bioinspired graphene membrane with temperature
Subsequently, the solvent was filtered. After adding 100 mL of pure tunable channels for water gating and molecular separation. Nat.
NMP to the cell to re-disperse the retentate, the solvent was filtered Commun. 8, 2011 (2017).
again. This process of solvent exchange was repeated once more. The 9. Kim, J. P. et al. Ultrafast H2-selective nanoporous multilayer gra-
solvent exchange process was conducted immediately after each phene membrane prepared by confined thermal annealing. Chem.
completion of solvent filtration. Commun. 57, 8730–8733 (2021).
10. Zhang, W. et al. General synthesis of ultrafine metal oxide/reduced
Computational methods graphene oxide nanocomposites for ultrahigh-flux nanofiltration
The molecular dynamics (MD) simulation approach was employed to membrane. Nat. Commun. 13, 471 (2022).
investigate the permeation behaviors of solvent molecules passing 11. Yang, Q. et al. Ultrathin graphene-based membrane with precise
through the graphene layers. First, each periodic simulation model was molecular sieving and ultrafast solvent permeation. Nat. Mater. 16,
initially prepared to contain graphene layers with solvent molecules 1198–1202 (2017).
occupying space between the layers. Either toluene, water, or ethanol 12. Qi, B. et al. Strict molecular sieving over electrodeposited 2D-
was introduced as a solvent in each simulation model. The prepared interspacing-narrowed graphene oxide membranes. Nat. Commun.
models were then equilibrated by a series of NVT (298 K) and NPT 8, 825 (2017).

Nature Communications | (2023)14:901 11


Article https://doi.org/10.1038/s41467-023-36524-x

13. Shen, Q. et al. Development of ultrathin polyamide nanofilm with 36. Dai, B. et al. High-quality single-layer graphene via reparative
enhanced inner-pore interconnectivity via graphene quantum dots- reduction of graphene oxide. Nano Res. 4, 434–439 (2011).
assembly intercalation for high-performance organic solvent 37. Cançado, L. G. et al. General equation for the determination of the
nanofiltration. J. Membr. Sci. 635, 119498 (2021). crystallite size La of nanographite by Raman spectroscopy. Appl.
14. Balaji, K. R. et al. Composite nanofiltration membrane comprising Phys. Lett. 88, 163106 (2006).
one-dimensional erdite, two-dimensional reduced graphene oxide, 38. Su, Y. S. & Manthiram, A. Lithium-sulphur batteries with a micro-
and silkworm pupae binder. Mater. Today Chem. 22, 100602 (2021). porous carbon paper as a bifunctional interlayer. Nat. Commun. 3,
15. Zhang, L. et al. Leaf-veins-inspired nickel phosphate nanotubes- 1166 (2012).
reduced graphene oxide composite membranes for ultrafast 39. Cao, Y. et al. New structural insights into densely assembled
organic solvent nanofiltration. J. Membr. Sci. 649, 120401 (2022). reduced graphene oxide membranes. Adv. Funct. Mater. 32,
16. Wang, X.-Y., Narita, A. & Müllen, K. Precision synthesis versus bulk- 2201535 (2022).
scale fabrication of graphenes. Nat. Rev. Chem. 2, 0100 (2018). 40. Konios, D., Stylianakis, M. M., Stratakis, E. & Kymakis, E. Dispersion
17. Kim, J. H. et al. Scalable fabrication of deoxygenated graphene behaviour of graphene oxide and reduced graphene oxide. J. Col-
oxide nanofiltration membrane by continuous slot-die coating. J. loid Interface Sci. 430, 108–112 (2014).
Membr. Sci. 612, 118454 (2020). 41. Huang, H. et al. Ultrafast viscous water flow through nanostrand-
18. Jang, J. et al. Turbostratic nanoporous carbon sheet membrane for channelled graphene oxide membranes. Nat. Commun. 4,
ultrafast and selective nanofiltration in viscous green solvents. J. 2979 (2013).
Mater. Chem. A 8, 8292–8299 (2020). 42. Wang, J. et al. A regularly channeled lamellar membrane for
19. Nie, L. et al. Realizing small-flake graphene oxide membranes for unparalleled water and organics permeation. Angew. Chem. Int. Ed.
ultrafast size-dependent organic solvent nanofiltration. Sci. Adv. 6, 57, 6814–6818 (2018).
eaaz9184 (2020). 43. Fu, W. et al. A high-flux organic solvent nanofiltration membrane
20. Cho, K. M. et al. Ultrafast-selective nanofiltration of an hybrid with binaphthol-based rigid-flexible microporous structures. J.
membrane comprising laminated reduced graphene oxide/gra- Mater. Chem. A 9, 7180–7189 (2021).
phene oxide nanoribbons. ACS Appl. Mater. Interfaces 11, 44. Liu, Q. et al. Molecular dynamics simulation of water-ethanol
27004–27010 (2019). separation through monolayer graphene oxide membranes: sig-
21. Zhou, Z. et al. Electropolymerization of robust conjugated micro- nificant role of O/C ratio and pore size. Sep. Purif. Technol. 224,
porous polymer membranes for rapid solvent transport and narrow 219–226 (2019).
molecular sieving. Nat. Commun. 11, 5323 (2020). 45. Foller, T. et al. Mass transport via in-plane nanopores in graphene
22. Kang, J. et al. Thermally-induced pore size tuning of multilayer oxide membranes. Nano Lett. 22, 4941–4948 (2022).
nanoporous graphene for organic solvent nanofiltration. J. Membr. 46. Kim, D. W. et al. Revealing the role of oxygen debris and functional
Sci. 637, 119620 (2021). groups on the water flux and molecular separation of graphene
23. Boukhvalov, D. W., Katsnelson, M. I. & Son, Y. W. Origin of anom- oxide membrane: a combined experimental and theoretical study.
alous water permeation through graphene oxide membrane. Nano J. Phys. Chem. C 122, 17507–17517 (2018).
Lett. 13, 3930–3935 (2013). 47. Lei, X., Tay, S. W., Ong, P. J. & Hong, L. Organic dye solution
24. Zheng, S., Tu, Q., Wang, M., Urban, J. J. & Mi, B. Correlating inter- nanofiltration by 2D Zn-TCPP(Fe) membrane - leverage of chemical
layer spacing and separation capability of graphene oxide mem- and fluid dynamic effects. J. Ind. Eng. Chem. 78, 410–420 (2019).
branes in organic solvents. ACS Nano 14, 6013–6023 (2020). 48. Kang, J. et al. Functionalized nanoporous graphene membrane with
25. Voiry, D. et al. High-quality graphene via microwave reduction of ultrafast and stable nanofiltration. J. Membr. Sci. 618, 118635 (2021).
solution-exfoliated graphene oxide. Science 353, 1413–1416 (2016). 49. Plimpton, S. Fast parallel algorithms for short-range molecular
26. Jang, J. H., Woo, J. Y., Lee, J. & Han, C. S. Ambivalent effect of dynamics. J. Comput. Phys. 117, 1–19 (1995).
thermal reduction in mass rejection through graphene oxide 50. Mayo, S. L., Olafson, B. D. & Goddard, W. A. DREIDING: a generic
membrane. Environ. Sci. Technol. 50, 10024–10030 (2016). force field for molecular simulations. J. Phys. Chem. 94,
27. Noh, S. H. et al. Joule heating-induced sp2-restoration in graphene 8897–8909 (1990).
fibers. Carbon 142, 230–237 (2019).
28. Kim, K. H. et al. The role of layer-controlled graphene for Acknowledgements
tunable microwave heating and its applications to the synthesis of This work was supported by the National Research Foundation of Korea
inorganic thin films. ACS Appl. Mater. Interfaces 8, 5556–5562 (NRF) grant funded by the Government of South Korea (MSIT) (NRF-
(2016). 2020R1C1C1003289), and the Technology Innovation Program
29. Hunter, C. A. & Sanders, J. K. The nature of. pi.-. pi. interactions. J. (20013621, Center for Super Critical Material Industrial Technology)
Am. Chem. Soc. 112, 5525–5534 (1990). funded by the Ministry of Trade, Industry & Energy (MOTIE, South Korea).
30. Cai, M., Thorpe, D., Adamson, D. H. & Schniepp, H. C. Methods of This research was supported by basic science research program
graphite exfoliation. J. Mater. Chem. 22, 24992–25002 (2012). through the National Research Foundation of Korea funded by the
31. Tu, Z. et al. Controllable growth of 1-7 layers of graphene by che- Ministry of Education (NRF-2019R1A6A1A11055660).
mical vapour deposition. Carbon 73, 252–258 (2014).
32. Wu, J. B., Lin, M. L., Cong, X., Liu, H. N. & Tan, P. H. Raman spec- Author contributions
troscopy of graphene-based materials and its applications in rela- J. Kang performed the experiments and collected the data. Y.K. and
ted devices. Chem. Soc. Rev. 47, 1822–1873 (2018). K.C.K. performed computational simulation. J.P.K., J.Y.K., J. Kim, and
33. Wollbrink, A. et al. Amorphous, turbostratic and crystalline carbon O.K. assisted in data analysis. J. Kang and D.W.K. prepared the manu-
membranes with hydrogen selectivity. Carbon 106, 93–105 (2016). script. The project was conducted under the supervision of D.W.K. All
34. Heller, E. J. et al. Theory of graphene raman scattering. ACS Nano subjects gave their informed consent for inclusion before they partici-
10, 2803–2818 (2016). pated in the study.
35. Shen, L. et al. Highly porous nanofiber-supported monolayer gra-
phene membranes for ultrafast organic solvent nanofiltration. Competing interests
Sci. Adv. 7, eabg6263 (2021). The authors declare no competing interests.

Nature Communications | (2023)14:901 12


Article https://doi.org/10.1038/s41467-023-36524-x

Additional information Open Access This article is licensed under a Creative Commons
Supplementary information The online version contains Attribution 4.0 International License, which permits use, sharing,
supplementary material available at adaptation, distribution and reproduction in any medium or format, as
https://doi.org/10.1038/s41467-023-36524-x. long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license, and indicate if
Correspondence and requests for materials should be addressed to Dae changes were made. The images or other third party material in this
Woo Kim. article are included in the article’s Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not
Peer review information Nature Communications thanks Suzana Nunes included in the article’s Creative Commons license and your intended
and the other, anonymous, reviewer(s) for their contribution to the peer use is not permitted by statutory regulation or exceeds the permitted
review of this work. use, you will need to obtain permission directly from the copyright
holder. To view a copy of this license, visit http://creativecommons.org/
Reprints and permissions information is available at licenses/by/4.0/.
http://www.nature.com/reprints
© The Author(s) 2023
Publisher’s note Springer Nature remains neutral with regard to jur-
isdictional claims in published maps and institutional affiliations.

Nature Communications | (2023)14:901 13

You might also like